Contrib Mineral Petrol DOI 10.1007/s00410-012-0758-0 ORIGINAL PAPER Multicomponent diffusion in garnets I: general theoretical considerations and experimental data for Fe–Mg systems Sascha André Borinski • Ulrich Hoppe • Sumit Chakraborty • Jibamitra Ganguly Santanu Kumar Bhowmik • Received: 2 November 2011 / Accepted: 12 April 2012 Ó Springer-Verlag 2012 Abstract We have carried out a combined theoretical and experimental study of multicomponent diffusion in garnets to address some unresolved issues and to better constrain the diffusion behavior of Fe and Mg in almandine–pyroperich garnets. We have (1) improved the convolution correction of concentration profiles measured using electron microprobes, (2) studied the effect of thermodynamic nonideality on diffusion and (3) explored the use of a mathematical error minimization routine (the Nelder-Mead downhill simplex method) compared to the visual fitting of concentration profiles used in earlier studies. We conclude that incorporation of thermodynamic non-ideality alters the shapes of calculated profiles, resulting in better fits to measured shapes, but retrieved diffusion coefficients do not differ from those retrieved using ideal models by more than a factor of 1.2 for most natural garnet compositions. Communicated by T. L. Grove. Electronic supplementary material The online version of this article (doi:10.1007/s00410-012-0758-0) contains supplementary material, which is available to authorized users. S. A. Borinski (&) S. Chakraborty Institut für Geologie, Mineralogie und Geophysik, Ruhr Universität Bochum, 44780 Bochum, Germany e-mail: s.borinski@gmx.de U. Hoppe Institut für Computational Engineering, Ruhr Universität Bochum, 44780 Bochum, Germany J. Ganguly Department of Geosciences, University of Arizona, Tucson, AZ 85721, USA S. K. Bhowmik Department of Geology and Geophysics, Indian Institute of Technology, Kharagpur 721 302, India Diffusion coefficients retrieved using the two kinds of models differ only significantly for some unusual Mg–Mn– Ca-rich garnets. We found that when one of the diffusion coefficients becomes much faster or slower than the rest, or when the diffusion couple has a composition that is dominated by one component ([75 %), then profile shapes become insensitive to one or more tracer diffusion coefficients. Visual fitting and numerical fitting using the NelderMead algorithm give identical results for idealized profile shapes, but for data with strong analytical noise or asymmetric profile shapes, visual fitting returns values closer to the known inputs. Finally, we have carried out four additional diffusion couple experiments (25–35 kbar, 1,260–1,400 °C) in a piston-cylinder apparatus using natural pyrope- and almandine-rich garnets. We have combined our results with a reanalysis of the profiles from Ganguly et al. (1998) using the tools developed in this work to obtain the following Arrhenius parameters in D = D0 exp{–[Q1bar ? (P–1)DV?]/RT} for D*Mg and D*Fe: Mg: Q1bar = 228.3 ± 20.3 kJ/mol, D0 = 2.72 (±4.52) 9 10-10 m2/s, Fe: Q1bar = 226.9 ± 18.6 kJ/mol, D0 = 1.64 (±2.54) 9 10-10 m2/s. DV? values were assumed to be the same as those obtained by Chakraborty and Ganguly (1992). Keywords Convolution effect Garnet Multicomponent diffusion Numerical model Thermodynamic non-ideality Introduction Compositionally zoned garnets occur in a wide variety of geological settings that include hydrothermal systems, skarns, acidic volcanics, igneous granites and their contact metamorphic aureoles, regional metamorphic pelites, 123 Contrib Mineral Petrol calc-silicates and metabasites, and ultramafic rocks from the mantle (see, e.g., Chakraborty and Ganguly 1991; Kohn 2003, for reviews). The preservation and modification of compositional zoning has been used widely to obtain timescales of various processes (e.g., cooling rates, exhumation rates) and as records of tectonic-, reaction-, deformation- and fluid-flow histories. A key parameter in such analyses is the knowledge of the relevant diffusion coefficients in garnets. A significant body of diffusion coefficients for different species in various garnet compositions is now available, and these have been recently reviewed by Ganguly (2010). Most experimental studies have focused on almandine– pyrope-rich garnet solid solutions that occur commonly in metamorphic rocks or in the mantle. Given the multicomponent nature of diffusion in garnet and the unavailability of suitable crystals to cover the compositional space adequately (e.g., nine crystals of well-defined composition to extract a 3 9 3 multicomponent diffusion matrix), the approach pioneered by Loomis et al. (1985) has been used by most authors (e.g., Chakraborty and Ganguly 1992; Ganguly et al. 1998; Perchuk et al. 2009) to determine diffusion coefficients from multicomponent exchange experiments with garnets. This approach makes use of the relationship between the different elements of the D-matrix via tracer diffusion coefficients. Measured concentration profiles generated by chemical diffusion (diffusion in response to a chemical concentration/chemical potential gradient) are fit to retrieve four tracer diffusion coefficients (D*Fe, D*Mg, D*Mn and D*Ca). We will refer to this approach as ‘‘multicomponent modeling’’ in the following. For applications, these tracer diffusion coefficients are then recombined using the same relationship to obtain D-matrices for various garnet compositions. The use of effective binary diffusion coefficients (EBDC) to model the multicomponent diffusion profiles (e.g., Chakraborty and Ganguly 1992; Vielzeuf et al. 2007) also relies implicitly on such connections. The manner in which the approach of Loomis et al. (1985) has been used so far relies, however, on several assumptions: (1) (2) (3) The model relating chemical diffusion coefficients (the elements Dij of the D-matrix in a multicomponent system) to tracer diffusion coefficients (D*i ) developed by Lasaga (1979) is valid. The effect of thermodynamic non-ideality has been ignored in all analyses so far, and the garnets have been assumed to be ideal solutions. A special case of the convolution correction of Ganguly et al. (1988) for systems with a single, constant diffusion coefficient was applicable to the measured concentration profiles in the multicomponent diffusion experiments using pyrope-almandine couple. 123 (4) The visual fitting of the concentration profiles to retrieve tracer diffusion coefficients provides true best fits in a statistical sense and is unique. Moreover, there remain issues as to the extent to which crystals used for the measurement of diffusion coefficients in the laboratory are representative of compositions found in the geological settings described above (e.g., Carlson 2002, 2006), and the parameters describing the diffusion coefficient of Ca remain poorly defined (Ganguly 2010). The first assumption was tested partially by Chakraborty and Rubie (1996) when they measured Mg tracer diffusion coefficients directly and found that the results agreed within error with extrapolation of data retrieved from multicomponent modeling. We have carried out new experiments and theoretical analysis to explore the validity of assumptions (2)–(4). We consider the experimental results and their implications for the diffusion of Fe and Mg in this paper; the diffusion behavior of Ca, which is found to be affected by added complications, is treated in a companion paper (referred henceforth as Part-II). The paper is organized as follows. First, we discuss the compositional space spanned by experimental data and how these relate to compositional dependence of diffusion coefficients and occurrences in natural systems. Next, we present the theoretical background for handling of experimental data that are developed in this paper. This section covers three aspects: a model for relating tracer diffusion coefficients to elements of the D-matrix in a 4-component non-ideal garnet solid solution, convolution correction of an arbitrary profile shape (i.e., without the assumption of single, constant diffusivity) and a statistical procedure for evaluating the fit between calculated and experimentally measured profiles in order to retrieve optimized values for the tracer diffusion coefficients. This is followed by a discussion of some general behavior of the diffusion profiles in this system and the sensitivity of profile shapes to values of different tracer diffusion coefficients. After this, the experimental methods and data are presented. The data from this study and other related works are then considered in order to retrieve Arrhenius parameters, and some implications of these results for applications to natural systems (e.g., closure temperatures) are discussed to conclude the paper. Compositional space of garnets in experiments versus nature Before embarking on a theoretical analysis of multicomponent diffusion behavior in garnets, it is worth evaluating to what extent the compositions of the garnet crystals used for laboratory measurements correspond to those in which Contrib Mineral Petrol compositional profiles is measured and interpreted in natural systems. Much of the discussion of compositional zoning of garnets focuses on the dodecahedral cations, and therefore, (Fe, Mg, Mn, Ca)3Al2Si3O12 is a suitable space to consider for this purpose. Considering ratios or the use of terms such as ‘‘almandine-rich’’ is ambiguous in that the concentration of the remaining dodecahedral cations remains inadequately specified in this description. To address this issue, we have plotted compositions from a number of metamorphic and mantle-derived garnets in a Fe–Mg–Mn–Ca tetrahedron, and in Ca–Mg–Fe and Mn–Mg–Fe ternaries and overlain the composition of diffusion couples used in laboratory experiments on these (Fig. 1). Several aspects emerge from this comparison: (1) A vast majority of metamorphic garnets are nearly binary almandine–pyrope solid solutions that are relatively poor in the spessartine and the grossularite components. Many experimental diffusion couples span this range exactly. (2) Low-grade garnets are often rich in the spessartine component. The compositional range of these garnets is well covered by the experimental diffusion couples of Chakraborty and Ganguly (1992). (3) Compositions of some mantle-derived and granulitic (particularly mafic granulites) garnets that are somewhat rich in the grossularite component are not well described by available experimental diffusion couples. However, most of these compositions are not very far removed from the compositions of many experimental diffusion couples. (4) Garnets from eclogitic rocks can be quite rich in the grossularite component, and these are not represented by any experimental diffusion couples. The implications of this comparison are that (a) the diffusion of low-grade, Mn-rich garnets occurs along different compositional vectors compared to that in most other, Fe–Mg-rich garnets. In the rest of this paper, we will refer to diffusion along these two compositional vectors as diffusion in ‘‘Mn-rich’’ and ‘‘Fe–Mg’’ garnets, respectively. It is implied that in all cases multicomponent diffusion of all four cations (Fe, Mg, Mn and Ca) occurs; the designation is merely an indication of the dominant compositional vector. (b) With the exception of certain eclogites, compositions of most natural metamorphic and mantle-derived garnets correspond very well to the range covered by experimental diffusion couples. Consideration of the diffusion couples along the ‘‘Fe–Mg’’ vectors also reveals that the compositional range spanned by a given couple can be quite variable. In some cases, the compositions of two ends of the diffusion couples are fairly close to each other so that the compositional vectors are short. Within this limited compositional range, it is possible to treat a diffusion coefficient as a constant and compositional profiles in such couples can be modeled adequately with a constant diffusion coefficient. On the other hand, there are couples that span almost the entire range of Fe–Mg compositions. For these couples, it is essential to consider the compositional dependence of diffusivity in modeling the compositional profiles (they often have a pronounced asymmetry in shape, e.g., as a result of such compositional dependence of diffusivity). This aspect will become important in the discussion that follows. Diffusion in ‘‘Mn-rich’’ garnet was studied by Chakraborty and Ganguly (1992) and Perchuk et al. (2009) only [an earlier study by Elphick et al. (1985) in the same laboratory used similar garnets, but those data were considered together in Chakraborty and Ganguly (1992)]. Most other diffusion experiments with multicomponent garnets dealt with ‘‘Fe–Mg’’ garnets. It is found that there is a qualitative difference between diffusion along these two kinds of compositional vectors/gradients. For example, in Mn-rich garnets, D*Fe*D*Mg, whereas in ‘‘Fe–Mg’’ garnets, D*Fe \ D*Mg (e.g., see Ganguly et al. 1998). Therefore, it is best to handle diffusion in these two systems separately. In this paper, we will focus on diffusion in ‘‘Fe–Mg’’ systems. Theoretical developments A multicomponent diffusion model for non-ideal garnet solid solutions According to the Lasaga (1979) model, the tracer diffusion coefficients and the elements, Dij, of the [D] matrix are related to each other by D zi zj Xi Dj Dn Dij ¼ Di dij Pn i 2 k¼1 Dk zk Xk ð1Þ in an ideal system and by .Xn o ln ci o ln ci zj Dij ¼ Di dij þ Di Xi D z2 X Di zi Xi k¼1 k k k oXj oXn zn " # n X o ln ck zi o ln ck zj D j D n þ X k z k Dk oXj oXn zn k¼1 ð2Þ in a non-ideal system, where dij = 0 if i = j and 1 if i = j. D*i , Xi, zi and ci represent the self-diffusion coefficient, molar fraction, valence and activity coefficient of component i, respectively. To use Eq. (2), it is necessary to have an activitycomposition model that allows one to calculate the (qlnci/ qXj) terms. For the garnet compositions of interest, such activity-composition relations have been determined experimentally by Ganguly et al. (1996), who treated the garnet solid solution according to a multicomponent solution model that was developed by Cheng and Ganguly (1994), which is based on a Taylor series expansion of 123 Contrib Mineral Petrol the multicomponent Gibbs free energy surface, as in the ternary formulation of Wohl (1953). This formulation has the attractive property that the compositions of the bounding binaries that are used to calculate the multicomponent free energy turn out, as a mathematical consequence, to be those obtained by the normal projections of the multicomponent composition on to the binaries. The normal projection method was suggested by the widely used Redlich–Kister–Muggianu formulation (Muggianu et al. 1975). Ignoring the ternary interactions, the general expression of ln ci in a multicomponent solution with subregular binaries is given, according to Cheng and Ganguly model, by " n X 1 ln ci ¼ ð1 2Xi Þ Wij Xj2 þ 2Xi ð1 Xi Þ RT j¼16¼i n X Wji Xj 2 j¼16¼i þ n n X X Wjk Xj Xk2 j¼16¼i k¼16¼i;j n1 X n X ð1 2Xi Þ Xj Xk ðWij þ Wik þ Wji 2 j¼16¼i k¼jþ16¼i þ Wjk þ Wki þ Wkj Þ n2 X Xj j¼16¼i n X n1 X Xk k¼jþ16¼i # Xl ðWjk þ Wjl þ Wkj þ Wkl þ Wlj þ Wlk Þ l¼kþ16¼i ð3Þ where Wij represents a subregular interaction parameter in the binary join i - j. The interaction parameters are independent of composition, but depend on pressure and temperature, and can be expressed as (Ganguly and Saxena 1987) Wij ðP; TÞ ¼ WijH ð1; TÞ TWijS ð1; TÞ þ ðP 1ÞWijV ; j¼16¼i;m Wjm Xj n1 X j¼16¼i;m Xj # þWjm þ Wjk þ Wkm þ Wkj Þ 123 n X " o ln ci 1 ¼ 2ð1 2Xi ÞWim Xm þ 2Xi ð1 Xi ÞWmi RT oXm n X 2 ð2Wjm Xj Xm þ Wmj Xj2 Þ j¼16¼i;m þ n ð1 2Xi Þ X Xj ðWim þ Wij þ Wmi þ Wmj þ Wji þ Wjm Þ 2 j¼16¼i;m n1 X j¼16¼i;m Xj n X # Xk ðWmj þ Wmk þ Wjm þ Wjk þ Wkm þ Wkj Þ k¼jþ16¼i;m ð5bÞ for all other cases. Figure 2 shows a set of compositional profiles calculated using ideal solution model, as in the earlier studies (Chakraborty and Ganguly 1992; Ganguly et al. 1998; Perchuk et al. 2009; Carlson 2006) those calculated using the same set of diffusion parameters, but with a non-ideal model using the interaction parameters in Table 1. The diffusion parameters and times for which the profiles are calculated correspond to parameters and experimental run durations that will be encountered below. Various pairs were chosen to study the effects and four extreme situations: (a) a diffusion couple close to the Fe–Mg join representing a composition from one of our experimental runs, (b) a diffusion couple lying close to the Fe–Ca join that might be found in an amphibole facies rock, (c) a diffusion couple where a high Fe- and a high Mn-rich composition, both including substantial Mg- and Ca-bearing components, are coupled—these represent compositions from certain granulitic rocks and cover a region of compositional space where some of the most non-ideal garnets may occur, and (d) a diffusion couple that spans the Mg-rich area of the tetrahedral space, but including substantial Mn- and Ca-bearing components, which stands for the composition ð4Þ S V where WH ij , Wij and Wij are, respectively, the enthalpic, entropic and volumetric components. In aluminosilicate garnets, values of these components are taken from Ganguly et al. (1996) and are reproduced here for convenience (Table 1). Using Eq. (3), (qlnci/qXj) terms required to calculate Dij according to Eq. (2) takes the form " n X o ln ci 1 2 ¼ Wmj Xj2 þ 2ð1 2Xm Þ RT oXm j¼16¼i;m n X when i = m and Xk ðWmj þ Wmk k¼jþ16¼i;m ð5aÞ Table 1 Summary of the internally consistent binary interaction parameters in aluminosilicate garnets after Table 4 in Ganguly et al. (1996) Parameter (ij) WH ij (J mol-1) WSij (J mol-1 K-1) WVij (J mol-1 bar-1) CaMg 21,627 5.78 0.012 MgCa 9,834 5.78 0.058 CaFe 873 1.69 0 FeCa 6,773 1.69 0.03 MgFe 2,117 0 0.07 FeMg 695 0 0 7.67 0.04 MgMn 12,083 MnMg 12,083 7.67 0.03 FeMn 539 0 0.04 MnFe 539 0 0.01 Contrib Mineral Petrol Fig. 1 Fe–Mg–Mn–Ca tetrahedron and ternary Fe–Mg–Ca and Fe– c Mg–Mn projections show typical compositions of garnets in garnetbearing rocks (triangles in tetrahedron, fields in ternary diagrams) and compositions of garnets used for experimental diffusion studies (diamonds). Each connecting line between the experimental sample compositions (diamonds) corresponds to one diffusion couple. Ternary projections additionally show diffusion couples used in Fig. 2 as circles with connecting lines. These lines represent hypothetical diffusion couples used for illustrating the effects of thermodynamic non-ideality on shapes of diffusion profiles in garnet (shown in Fig. 2). Data used for drawing this figure are available online as electronic supplementary material of garnets in certain mantle xenoliths (see Fig. 1). In all of the four situations, compositions with a relatively high concentration of Ca were considered, that is, where a strong non-ideality effect is expected. In addition, in the last two cases, high concentrations of (Mn and Mg) were also present on at least one side of the diffusion couple. This is an unusual situation, but one where effects of nonideality are expected to be maximized because of the high values of the Mg–Mn interaction parameter (Table 1). The difference between the profile shapes calculated using the ideal and the non-ideal formalisms is found to be minimal in the first three cases (*1 mol. % at any point along the profile). This is consistent with the expectation based on a quasi-binary approximation of the mixing property (Chakraborty and Ganguly 1992). The diffusion coefficients that would best describe a given profile using either the ideal or the non-ideal model lie within factor of 1.2 of each other. An exception is given in the fourth case where the Mg, Ca and Mn concentrations are higher. Here, the difference between the shapes obtained using the two formalisms may result in differences of up to *4 mol. % at some spatial coordinates, and the diffusion coefficients lie within a factor of 2.5. Such a difference is significant and is an example where non-ideality cannot be neglected in these compositionally unusual situations, since only a rare variety of mantle-derived garnet have relatively high Mn content (6 mol. %) in addition to high Mg and Ca. This is a case where the effect of non-ideality is maximized because of strong Mg–Mn and Mg–Ca interactions. In all cases, the diffusion coefficients required to describe a given profile using a non-ideal model will be greater. What is noticeable, however, is that although profile lengths are not affected substantially, the profile shapes are modified and some inflections/changes of curvature (including effects of uphill diffusion) that are obtained with a non-ideal model cannot be produced using an ideal model (Fig. 2). Moreover, the effects of non-ideality are different for different elements, and the symmetry of profile shapes changes when non-ideal models are considered. For example, the point of inflection of a profile can be shifted from the central segment around the interface of the diffusion couple, creating the 123 Contrib Mineral Petrol a1 a2 b1 Mg Fe 0.70 0.03 b2 0.03 0.70 Mn Ca -0.01 0.30 0.20 0.60 Ca Mg 0.50 0.01 0.40 Mn -0.01 0.30 Mn 0.20 0.10 -1 0 1 2 -2 2 0 -2 -1 Distance [µm] c1 c2 Ca 0.01 0.40 Mn -0.01 Fe 0.20 N o r m . C o n c e n tr a tio n Mg 0.50 Cnon-ideal - C -ideal Mn Mg 2 -2 Mg Mg 0.03 0.60 Fe 0.50 0.01 0.40 Fe Mn -0.01 0.30 0.20 Ca Ca 0.10 -0.03 -2 -1 0 1 2 2 0 d2 0.70 0.60 Ca 1 d1 0.03 0.30 0 Distance [µm] Fe 0.70 -0.03 -2 0 2 Distance [µm] 0.10 -0.03 Mn -2 -1 Cnon-ideal - C -ideal -2 Mg 0.10 -0.03 Mn Ca N o r m . C o n c e n t r a t io n Fe Ca Cnon-ideal - C -ideal 0.40 0.01 N o r m . C o n c e n tr a tio n Mg Fe 0.50 Cnon-ideal - C -ideal N o r m . C o n c e n tr a tio n Fe 0.60 0 1 2 -2 0 2 Distance [µm] Fig. 2 a1–d1 Four sets of simulated diffusion profiles with arbitrarily chosen initial composition differences. All profiles have been simulated with the thermodynamically ideal (dashed lines) and nonideal solution (solid lines) models using a constant D-matrix. All profiles are calculated using the following set of diffusion coefficients (Mg 1.03 9 10-19, Fe 4.00 9 10-20, Mn 3.79 9 10-19, Ca 2.46 9 10-20 m2/s). These diffusion coefficients are representative for 28 kbar, 1,050 °C. Profile lengths correspond to an annealing time of 440 h. Note that D*-values and run conditions used here are equivalent to the experimental run DPA4. Vertical dashed lines represent initial point of contact between garnets of the diffusion couple. a2–d2 Difference between concentration profiles calculated using ideal and non-ideal solution models, respectively appearance of a Kirkendall shift such as those known in metallic systems (Smigelskas and Kirkendall 1947). Therefore, statistical measures of quality of fit are likely to be improved when a non-ideal model is used, and we have used this model (Eq. 2) for analyzing the data discussed in this paper. 1998), it is necessary to correct for the convolution that arises from the spatial averaging effect of the microprobe beam. Ganguly et al. (1988) developed an analytical formulation to address this problem, assuming that the excitation volume for a spot analysis has Gaussian distribution of intensity with radial symmetry about the axis of the electron beam. They also developed an analytical solution in which the diffusion coefficient determined from a measured diffusion profile, which conforms to the solution of the diffusion equation with constant diffusion coefficient, is directly corrected for the convolution effect in the microprobe measurement of the profile. The second method was applied by Ganguly et al. (1998) to deconvolve Convolution analysis It has been shown that for the slow diffusion rates and resulting short profile lengths (often \10 lm) encountered in experimental studies of diffusion in garnet (e.g., Elphick et al. 1985; Chakraborty and Ganguly 1992; Ganguly et al. 123 Contrib Mineral Petrol a 0.70 Fe 0.60 N o rm . C o n c . the tracer diffusion coefficients retrieved from modeling multicomponent diffusion profiles in the pyrope–almandine couples. However, this procedure is not entirely satisfactory because the tracer diffusion coefficient of a species is not solely responsible for its diffusion profile in a multicomponent system. Hence, it is desirable to adopt the more general deconvolution analysis. Since we use a numerical finite difference scheme to calculate diffusion profiles, it is best to simply evaluate the convolution integral directly without any further approximations. This is done using Ca 0.50 0.40 0.30 Mn 0.20 Mg 0.10 fi ðxÞ ¼ðue Xi ÞðxÞ Zþ1 ¼ ue ðn xÞXi ðnÞdn -3 ð6Þ -2 -1 0 1 2 3 Distance [µm] 1 in which e represents the error standard deviation of the Gaussian or convolution factor. In our simulations of experimental diffusion profiles, we calculate profiles of all four components with an assumed set of (tracer) diffusion coefficients for the given run duration and then convolve the profiles using Eq. (6). This convolved profile is then compared to the experimental profiles determined by sequential spot analyses along a line traverse in an electron microprobe. The convolution factor (e = 0.48 lm) was determined earlier by Ganguly et al. (1998) by comparing profiles across a natural garnet–garnet diffusion couple measured by electron microprobe and analytical TEM, assuming that the convolution effect in the latter was negligible. Objective fits to experimental diffusion profiles One of the concerns, in spite of the repeated use of the visual fitting technique introduced by Loomis et al. (1985), is that an objective measure of the quality of fit has not been used in the method. Further, it has remained open to question how unique the retrieved diffusion coefficients are (can a given profile be described equally well by two different sets of diffusion coefficients?) and how sensitive the profile shapes are to small perturbations in the values of the diffusion coefficients. We have now developed a method to objectively test these aspects. The method is based on the Nelder-Mead algorithm or simplex search algorithm (Nelder and Mead 1965), which is one of the best-known algorithms for multidimensional unconstrained optimization without derivatives. The method is widely used to b 0.70 0.60 N o rm . C o n c . where fi(x) is the observed concentration and Xi(x) is the true concentration of component i at a given point of the diffusion profile. ue defines the so-called standard normal (or Gaussian) density function 1 1 x2 uðxÞ ¼ pffiffiffiffiffiffi exp ð7Þ 2 e e 2p Fe Ca 0.50 0.40 0.30 Mn 0.20 Mg 0.10 -3 -2 -1 0 1 2 3 Distance [µm] Fig. 3 Two sets of simulated diffusion profiles using the NelderMead optimization algorithm with an arbitrarily chosen initial composition. The profiles have been calculated with the same diffusion coefficients and simulation conditions as given in Fig. 2. a In the first set of simulations, the calculated diffusion profiles have been fitted to obtain diffusion coefficients that agree with the true values within three decimal places and the best-fit profiles are indistinguishable from the input (i.e., what would be measured) profiles. b Diffusion coefficients retrieved in the second set of simulations, where a bit of random noise was added to the profiles, have a small, but measurable deviation from the real diffusion coefficients even though the best fit and input (i.e., measured) profiles match very well. Diffusion coefficients differ by a factor of 1.01 for Mg, 1.09 for Fe, 0.98 for Ca and 1.03 for Mn solve parameter estimation and similar statistical problems, where the function values are uncertain or subject to noise. By progressively moving away from the poorest value, it locates the minimum in the difference between an objective function and measured values in a many dimensional space (e.g., the space of the four tracer diffusion coefficients, in our case). The validity of the method is demonstrated (Fig. 3) by calculating diffusion profiles with a given set of diffusion coefficients and then using arbitrary starting seed values to test whether the ‘‘true’’ diffusion coefficients are retrieved by the optimization program. Profiles calculated 123 Contrib Mineral Petrol according to the retrieved ‘‘best-fit’’ diffusion coefficients are shown for comparison in Fig. 3. In a second set of simulations, random noise was added to the calculated profiles to simulate analytical error and other sources of uncertainty in the measurement of compositions. For a range of values, agreements between the retrieved and true diffusion coefficients are excellent, with a maximum recorded difference of a factor of 1.09. One of the issues that have remained unaddressed explicitly is, to what extent is it possible to retrieve four tracer diffusion coefficients from one set of diffusion profiles? The optimization algorithm tests this aspect as well and demonstrates that, depending on the region of compositional space, it is possible to retrieve four tracer diffusion coefficients from a single diffusion couple experiment. We explore this aspect further below. We note here that in the absence of a known function that defines the compositional dependence of tracer diffusion coefficients, it is not possible to determine best-fit diffusion coefficients from compositional profiles that are visibly strongly asymmetric, requiring compositionally dependent tracer diffusion coefficients to describe their shapes. We have used the algorithm to retrieve diffusion coefficients from experimental diffusion couples where the compositional range spanned by a diffusion couple is not large (e.g., the red compositional vectors in Fig. 1) and the measured profiles are reasonably symmetric. This is discussed below after we present the results from the new diffusion experiments. Model calculations to test sensitivity and robustness To explore how sensitive the profile shapes are to the values of diffusion coefficients and how uniquely the shapes of these profiles are related to the different tracer diffusion coefficients, we have carried out a number of numerical simulations. In our exploration of compositional space, we have restricted ourselves to the range found to be relevant for the garnet compositions commonly found in nature (Fig. 1). The overall nature of the two effects can be seen by inspecting the equation relating binary inter-diffusion in an ideal, ionic system of equally charged species (Lasaga 1979): Dij ¼ Di Dj Di Xi þ Dj ð1 Xi Þ ð8Þ when all D*i s are similar to each other (as, e.g., is likely for the diffusion of similarly charged cations on a given lattice site of a particular mineral), then as the content of one component becomes very dilute, Dij*D*i . The magnitude of D*j plays no significant role in determining the profile 123 shapes, and conversely, D*j cannot be determined by fitting measured profiles. For multicomponent garnets with compositions relevant for experimental and natural samples, we have carried out a series of simulations to explore the compositional limit where this effect appears. It is found that when the concentration of any one component exceeds 75 mol. %, profile shapes of all components begin to get insensitive to the tracer diffusivity of that element, that is, the tracer diffusivity of that element can be poorly constrained from measured profile shapes. An example for four components is shown in Fig. 4. This suggests that diffusion couples made of intermediate garnet compositions yield the most well-constrained diffusion data. The second effect arises when one of the D*i s values differs significantly from that of all others. This effect can be explored using the binary equation (Eq. 8) as well. When D*i D*j , then Dij*D*j /Xi. Simulations in the multicomponent systems show that when the D* value for any one element is more than four orders of magnitude faster than that of all others, the profile shapes become insensitive to further changes in D* of this element. For example, the profile shapes for all elements are practically identical when D*A [ 103 (D*B, D*C and D*D), irrespective of the value of D*A (Fig. 4). The sensitivity of the profile shape to changes in values of D*A decreases as the difference between D*A, and the other tracer diffusivities increases. Thus, the method is not very suitable for the determination of diffusion coefficients where one of the diffusivities is much faster (several orders of magnitude) than the rest. However, this is not a common problem when diffusion of ions of the same charge on the same lattice site of a given mineral is considered (see, however, the discussion for Ca in Part-II). These aspects have important implications for multicomponent modeling of diffusion profiles in natural garnets to extract timescales (cooling rates, exhumation rates, or duration of metamorphism) as well. Experimental methods Four new experiments were carried out in Bochum in a 1,000 ton end-loaded piston-cylinder apparatus using WC cores with 12.7 mm internal diameters and WC pistons and talc-glass pressure cells with anhydrous parts within a graphite internal resistance furnace. Parts inside the graphite furnace of the pressure cell were dried to ensure that the fO2 was defined very closely by graphite–O2 equilibrium. Generally, the basic experimental approach is in practice the same as that used in the earlier studies from the Arizona group (Elphick et al. 1985; Chakraborty and Ganguly 1992; Ganguly et al. 1998), where a schematic Contrib Mineral Petrol 0.80 0.80 a Mg 0.70 Fe Fe Fe Mg Mg 0.60 N o rm . C o n c . 0.60 N o rm . C o n c . b 0.70 0.50 0.40 0.30 Mn 0.20 0.50 0.40 0.30 Mn 0.20 Ca Ca 0.10 0.10 0.00 1 2 3 4 5 6 0.00 7 1 2 Distance [µm] 0.80 4 5 6 0.80 c Mn d Ca 0.70 Fe Fe Mg Mg 0.60 N o rm . C o n c . 0.60 0.50 0.40 0.30 Mn 0.20 0.50 0.40 0.30 Mn 0.20 Ca Ca 0.10 0.00 7 Distance [µm] 0.70 N o rm . C o n c. 3 0.10 1 2 3 4 5 6 7 Distance [µm] 0.00 1 2 3 4 5 6 7 Distance [µm] Fig. 4 One set of calculated diffusion profiles (bold lines). Initial compositions, profile lengths and simulation conditions are equivalent to experimental run (DPA4) conducted in Ganguly et al. (1998). Fine lines: maximal adjustment that concentration profiles undergo when D*-value of one component (in a for Mg, b for Fe, c for Mn, d for Ca) is increased to approximately three orders of magnitude, while the other D*-values are held constant, so that D*i *[103 D*j=i. Higher D*-value of that component would not cause any further change in the calculated profile shape of any component. Therefore, the gray fields between bold and fine lines represent the sensitivity range—variation of profile shapes within these limits would allow a D*-value for any one of the components to be determined (see case 2 in text). Sensitivity range (to change of D*-values) of a component becomes smaller when that component begins to exceed 75 mol. % (see case 1 in text) cross section of the pressure cell and a more detailed description of the experimental procedure are available. Diffusion couples were made of gem-quality single crystals of natural Fe–Mg garnets where the couple forms a slightly conic-shaped cylinder with a vertical interface. The geometry of diffusion couples reduces uniaxial stress during cold compression. Before encasing the highly polished crystals in the graphite capsule, crystals were cleaned in an ultrasonic bath successively with acetone, alcohol, soapy water and deionized water to ensure that the crystals are in clean contact at the interface. The graphite capsule was then surrounded by a molybdenum tube of 0.2 mm wall thickness where the molybdenum tube serves to accomplish two purposes: (1) It reduces the thermal gradient within the tube in the vertical direction due to excellent heat conductivity. (2) It ensures that interfaces stay together under hot conditions due to the higher thermal expansion coefficient of the metal compared to the graphite-garnet assembly. Temperature was measured using a type-D thermocouple (W 25 %Re–W 3 %Re), and pressure was calibrated at high temperature using the quartz–coesite phase transformation. Pressure was increased manually over a period of 1 h, and then, the samples were annealed under computer control with a heating rate of 30 °C/min. After the expiration of experimental duration, the samples were quenched instantaneously and then decompressed slowly. To prepare the specimen for electron microprobe analysis, the diffusion couple was recovered from a high P–T experiment, sectioned and impregnated with low viscosity epoxy. Each section was carefully oriented within the epoxy and then polished so that the surface exposed to electron microprobe analyses was normal to the interface. Diffusion profiles of all cations present in the natural garnet samples were measured by step or beam scanning in an electron microprobe (Cameca SX-50) along a line normal to the interface. Composition of diffusion couples (see Fig. 1), experimental run conditions and results from fitting the diffusion profiles are reported in Table 2. 123 123 Pyr21Alm70Sps4.5Grs4.5- Pyr51Alm37Sps1Grs11 Pyr12Alm77Sps1Grs10- Pyr80Alm20Sps0Grs1 Pyr5Alm73Sps21Grs1- Pyr51Alm37Sps1Grs11 Pyr21Alm70Sps4.5Grs4.5- Pyr51Alm37Sps1Grs11 Pyr28Alm69Sps1Grs2- Pyr47.6Alm46.7Sps1Grs12 Pyr5Alm73Sps21Grs1- Pyr51Alm37Sps1Grs11 Pyr28Alm69Sps1Grs2- Pyr47.6Alm46.7Sps1Grs12 DPA7a GD13a DPA5a DPA10a R11a DPA 6a R17a 38,000 38,000 40,000 38,000 30,000 20,000 26,000 28,000 35,000 1,432 1,375 1,344 1,280 1,250 1,200 1,100 1,057 1,400 1,320 1,300 1,260 T (°C) 94.8 73.5 72.8 218 193 168 576 438 48 120 120 200 t (h) Ca 842 (386) 361 (178) 281 (106) 88 (31) 195 (66) 113 (29) 15 (7) 10 (5.3) 925 (246) 169 (39) 262 (95) 135 (40) 336 (144) 249 (102) 278 (106) 64 (24) 78 (24) 77 (22) 10 (5) 4 (1.4) 189 (64) 197 (44) 205 (78) 85 (26) 76 (50) 31 (18) 16 (13) 38 (18) 238 (62) 118 (140) 74 (40) 165 (110) 12 (10) 9.5 (5) 26 (18) 0.2 (0.2) 2.5 (2.2) 124 (71) 1,000 (157) 550 (147) 400 (116) 200 (68) 50 (17) 150 (34) 15 (5.3) 10 (4.7) 620 (202) 300 (68) 240 (63) 220 (65) Mg Mn Mg Fe D* (V) D* (NM) 1020 D* (2r) (m2/s) 480 (75) 200 (53) 280 (81) 200 (70) 40 (12) 80 (17) 13 (5.2) 7 (3.9) 540 (176) 190 (43) 150 (30) 95 (28) Fe 120 (79) 50 (18) 20 (17) 11 (5.1) 270 (69) Mn 100 (34) 50 (36) 200 (105) 5 (2.8) 20 (12) 20 (9.1) 0.8 (0.6) 0.5 (0.4) 150 (34) Ca 0.84 0.66 0.70 0.44 3.90 0.75 1.00 1.03 1.50 0.56 1.09 0.62 Mg 0.70 1.25 1.00 0.32 1.95 0.97 0.73 0.57 0.35 1.03 1.36 0.9 Fe 0.63 0.61 0.78 3.45 0.88 Mn D* (NM)/D* (V) 1.18 1.48 0.82 2.49 0.47 1.29 0.24 4.92 0.82 Ca a Results from our adjusted model fits to experimental data reported in Ganguly et al. (1998) Two sets of diffusion coefficients are shown—those fitted by using the Nelder-Mead (NM) algorithm and ‘‘visually’’ (V), respectively. All results are corrected for the convolution effect of the microprobe beam. ±2r errors for diffusion coefficients fitted by (V)-method are estimated using the accuracy of the distance measurement. Error of (NM) method additionally considers the error arising from the scatter of the data Pyr43Alm44Sps2Grs11- Pyr48Alm39Sps2Grs11 Pyr5Alm73Sps21Grs1- Pyr51Alm37Sps1Grs11 DPA4a 25,000 Pyr53Alm33Sps1Grs9Adr3Pyr74Alm13Sps1Grs4Adr3Uv4 SMG6 SMG0 25,000 Pyr52Alm39Sps4Grs4Adr1Pyr73Alm15Sps1Grs5Adr3Uvr4 SMG1 25,000 Pyr60Alm28Sps1Grs11Pyr74Alm14Sps1Grs5Adr3Uvr4 SMG5 P (bar) Diffusion couple Run no. Table 2 Summary of annealing conditions and self-diffusion data of divalent cations in pyrope–almandine diffusion couples Contrib Mineral Petrol Contrib Mineral Petrol Determination of diffusion coefficients Diffusion coefficients were retrieved from the measured concentration profiles using two different approaches. As a first approach, the minimization of the L1-distance between the measured and the computed profile according to the Nelder-Mead algorithm outlined above was used. The ‘‘visual fitting’’ approach of Loomis et al. (1985) was used to carry out forward modeling numerically, where diffusion profiles were calculated for chosen values of the four tracer diffusion coefficients (D*Fe, D*Mg, D*Mn and D*Ca), convolved using the approach outlined above and compared to the measured profiles (Fig. 5). The chosen diffusion coefficients were varied until the best description of the measured profile shapes was obtained. The input D* values used to obtain such a profile were considered to be the relevant tracer diffusion coefficients at that particular run condition. Diffusion coefficients retrieved according to both methods are reported in Table 2. The statistical errors for the tabulated values of self-diffusion coefficients derived from the electron microprobe data have been calculated as follows, taking into account the scatter of the data and resolution of step scan and/or beam scan profiling, which produces an uncertainty of length of the diffusion profile due to a stretching or contraction of the distance scale (&0.5 lm). Thus, the variance of a D* value, r2*D, including these errors, may be approximated as (Sano et al. 2011): rD r2D ðsÞ þ r2D ðxÞ where r2*D(s) and r2*D(x) are, respectively, the variance of D* arising from the two sources of errors described above. The first term considers the scatter of the data and has been estimated from the error of the optimization procedure by Nelder-Mead only. This term has not been taken into consideration for ‘‘visual fitting’’ of profiles. Following Sano et al. (2011), the second term caused by the uncertainty of diffusion distance is approximated as " # ldp þ rx 2 rD ðxÞ D 1 ð10Þ ldp a Interface Diffusion profile Prp52 Prp73 50.00 µm 0.80 BSE 15.0 kV 500 b 0.70 Mg N o rm . C o n c. 0.60 0.50 0.40 Fe 0.30 0.20 Ca 0.10 Mn 0.00 1 2 3 4 5 ð9Þ 6 7 Distance [µm] Fig. 5 a Back-scattered electron image of a diffusion couple (run SMG 1) annealed at 25 kbar, 1,300 °C for 120 h in a graphite capsule. Dark region represents the pyrope richer garnet crystal. Line across the interface represents the diffusion profile measured along a crack free region. b Corresponding Mg, Fe, Mn and Ca diffusion profiles. Lines through data points represent convolution corrected simulations, which were fitted by the following set of diffusion coefficients (Mg 2.62 9 10-18, Fe 2.05 9 10-18, Mn 2.38 9 10-18, Ca 1.24 9 10-18 m2/s) where ldp is the best estimate of the total diffusion distance and rx the estimated error of ldp. Note that other important sources of errors that may be present (e.g., from the reproducibility of experimental run conditions, or the use of garnets of somewhat different compositions in the diffusion couple experiments) are not included in these values. It is found that the D*Mg [ D*Fe at any given run condition, which is consistent with the findings in Fe–Mg systems in earlier studies (Elphick et al. 1985; Ganguly et al. 1998; Perchuk et al. 2009). D*Mn is poorly constrained (see also the discussion above), and D*Ca requires special handling—this is discussed in the companion paper. In order to obtain the best possible constraints on Arrhenius parameters, we have used data from diffusion profiles from all experiments carried out with Fe–Mg diffusion couples in piston-cylinder apparatus using a similar setup. The profiles were refitted using the improved techniques developed in the paper. We have shown that both techniques, (a) Nelder-Mead optimization and (b) ‘‘visual fitting’’, yield reliable diffusion coefficients and differ for Mg and Fe by a factor of *3 at most (Table 2). However, in order to have a set of most reliable diffusion coefficients, we have used the first approach for our final data analysis (Fig. 6). 123 Contrib Mineral Petrol 1400 1200 1300 term in the above expression represents an effective activation volume that incorporates the pressure effect of variation of fO2 along the reaction curve as well (Ganguly et al. 1998; Holzapfel et al. 2007). Here, we have assumed, based on results from earlier studies (e.g., Ganguly et al. 1998), that the activation volume of diffusion is not strongly variable for different garnet compositions and so we have taken the values of DV? from Chakraborty and Ganguly (1992) to be applicable to the Fe–Mg garnets as -1 well. These are DV? mol-1, DV? Mg = 0.53 J bar Fe = 0.56 -1 -1 J bar mol . With these activation volumes, fits to the diffusion coefficients retrieved from the experimental data yield the following values for the parameters in Eq. (11): 1100 °C -16.5 -17.0 Mg 2 l o g D ( m /s ) -17.5 -18.0 -18.5 Mg: Q1bar = 228.3 ± 20.3 kJ/mol, D0 = (2.72 ± 4.52) 9 10-10 m2/s, Fe: Q1bar = 226.9 ± 18.6 kJ/mol, D0 = (1.64 ± 2.54) 9 10-10 m2/s. -19.0 Fe -19.5 6.00 6.25 6.50 6.75 7.00 7.25 7.50 4 10 /T(K) Fig. 6 Arrhenius plot of experimentally determined self-diffusion coefficients of Mg (diamonds) and Fe (squares) in garnet fitted by the ‘‘visual’’ method. The error bars on diffusion data indicate approximately ±2r estimates (95 % confidence interval) of the individual D*-values, when only errors of measurement of composition and distance are considered (see text for more details). Solid lines are resulting from weighted least squares fits, whereas dotted lines represent the 2r envelopes of the log D*-values predicted from the Arrhenius relations and 2r errors of the Arrhenius parameters. All data have been normalized to a pressure of 25 kbar and fO2 corresponding to those defined by the presence of graphite in the system C–O–H. Filled symbols: diffusion data determined as part of this experimental study, open symbols: diffusion data determined in this study by refitting the experimental profiles of Ganguly et al. (1998) Given the limited number of data points that are available even after the repeated efforts at measuring these experimentally demanding diffusion coefficients, it is prudent to include as many data sets as possible in order to get meaningful Arrhenius parameters. For older data, it is also indicated where the data were reported for the first time and when the diffusion coefficients obtained in this study by refitting the earlier profiles are different from the values reported in the original papers. Assuming the activation volume, DV?, to be independent of pressure, the diffusion coefficient of a species as a function of pressure and temperature is given by Q1bar þ DV þ ðP 1Þ 2:303RT QP : ¼ log D0 2:303RT log D ðP; TÞ ¼ log D0 ð11Þ The diffusion coefficients in this study are determined from experiments at fO2 conditions along a curve defined by graphite–oxygen–hydrogen equilibrium in the pressure cell (imposed by the graphite capsule). Therefore, the DV? 123 Arrhenius parameters were calculated using the method of least squares, and Arrhenius slopes for Mg and Fe are shown in Fig. 6. Their ±2r error envelopes were determined according to Tirone et al. (2005) as 2 1 2 2 rlog D ðTÞ ¼ rlog D0 þ r T QP=2:303R 2 QP þ cov log D0 ; ð12Þ T 2:303R where r is the standard deviation, cov(logD0, Qp/2.303R) is the covariance of intercept (logD0) and slope (Qp/ 2.303R) (e.g., see Tirone et al. 2005). Uncertainties of results presented here contain only errors from the sources described above. The new results, obtained using a different population of natural garnets from the ones used in the study of Ganguly et al. (1998), demonstrate that (a) for the given range of Fe–Mg solid solution, the relatively more Mg-rich diffusion couples spanning a smaller compositional range (and hence more amenable to visual as well as numerical fitting) yield comparable diffusion coefficients, (b) differences in minor and trace elements between garnets do not have a significant effect on diffusion rates of elements such as Fe and Mg. Discussion Diffusion coefficients (D*Mg and D*Fe) obtained in this study are compared to results from earlier studies in Fig. 7. The consequences for modeling processes in metamorphic garnets using the results from this study may be explored by extreme extrapolation from the experimental run conditions to a P and T of 5 kbar and 600 °C, respectively. Tracer diffusion coefficients that would be obtained using the expressions given above can be compared to Contrib Mineral Petrol a (Cygan and Lasaga 1985), 3.0(Chakraborty and Rubie 1996), 242 (Schwandt et al. 1995), 2.1 (Ganguly et al. 1998), 1.5 (Carlson 2006) and 10.4 (Perchuk et al. 2009). For Fe tracer diffusion, h = 32 (Ganguly et al. 1998), 2.6 (Carlson 2006) and 1.9 (Perchuk et al. 2009). At this pressure condition, results from this study indicate that D*Mg/D*Fe lie within a factor of *1.4 between 600 and 1,400 °C, or in other words, the results obtained from additional experiments and improved fits in this study lie within a factor of 1.0–3.0 of earlier results, with the exceptions of D*Mg from Schwandt et al. (1995), D*Fe from Ganguly et al. (1998) and D*Mg from Perchuk et al. (2009). 1/h of the D*Mg is the approximate factor in each case by which timescales retrieved using the earlier results would differ from those that would be obtained using the present data set. Implications for resetting and closure of garnets in natural settings b Fig. 7 Arrhenius plot of (a) Mg and (b) Fe tracer/tracer diffusion coefficients determined in this study (solid bold lines) with earlier data. All data have been normalized to a pressure of 25 kbar and fO2 corresponding to those defined by graphite in the system C–O–H. Abbreviations: C06 Carlson (2006) obtained from earlier experimental data and modeling result of natural stranded diffusion profiles, CG92 (Chakraborty and Ganguly 1992) & G98 (Ganguly et al. 1998). D*-values retrieved from experimental data in Alm-Sps and Alm-Prp diffusion couples, respectively; CL85 (Cygan and Lasaga 1985), CR96 (Chakraborty and Rubie 1996) & S95 (Schwandt et al. 1995). D*(Mg) in natural Prp garnets, L85 (Loomis et al. 1985) (these data were incorporated in G98), P09 (Perchuk et al. 2009) Diffusion data in garnet find application in several situations. In addition to the modeling of frozen compositional profiles for geospeedometry, two common and related questions are: (i) For what kind of thermal histories can garnets of a given size retain their chemical compositions at the core? (ii) During cooling, at what temperatures do compositions at the cores of garnets of a given size freeze in? These are issues related to the concept of a closure temperature that are usually evaluated using the formulation developed by Dodson (1973). This formulation is, however, based on certain simplifying assumptions. One of these is that the compositional zoning in the mineral of interest should not retain any memory of the initial concentration distribution, that is, even at the core of the crystal the initial concentration should be reset by diffusion. This is why there is no term accounting for the initial concentration distribution or peak temperature in the formulations of Dodson. In garnets, however, this is frequently not the case; more often than not, diffusional zoning does not penetrate to the core. For such situations, Ganguly and Tirone (1999) have developed a formulation where they showed that the core composition of a spherical grain surrounded by a homogeneous infinite matrix (e.g., garnet surrounded by a large mass of biotite or a matrix with fast grain boundary diffusion) is not affected if the value of a dimensionless variable, M, is B0.1 for thermal histories with peak temperature, T0 B 1,000 °C. The parameter M is defined as M¼ results using expressions given in other studies on Fe–Mg garnets in terms of a factor h, where h[D*i from study X] = [D*i from this study]. For Mg tracer diffusion, h = 1.0 RDðT0 ÞT 2 QðdT=dtÞa2 ð13Þ where D(T0) is the diffusion coefficient at the peak temperature, a is a characteristic grain dimension (radius for a 123 Contrib Mineral Petrol sphere and infinite cylinder and half-thickness for an infinite plane sheet) and dT/dt is the cooling rate at T (the temperature that appears in the numerator). The cooling curve is assumed to follow an asymptotic relation in which the reciprocal temperature varies linearly with time (1/T = 1/T0 ? gt) (it is formally the same as that used by Dodson (1973) in his derivation of closure temperature formulation). Using Eq. (13), Tirone and Ganguly (2010) addressed the problem of resetting of core composition of garnet grains as a function of peak temperature, T0, and cooling rate. Following their approach, we illustrate (Fig. 8, similar to Fig. 2 of Tirone and Ganguly 2010) minimum grain sizes for which garnets would retain their core compositions (for a binary chemical diffusion coefficient, D(Fe–Mg) at XFe = 0.75 and XMg = 0.25, which is taken to be a typical composition for metapelitic garnet). The D(Fe–Mg) values can be calculated from Eq. 8, using the D*Fe and D*Mg values from this work. The results in Fig. 8 show that the minimum grain size to preserve core composition is revised upwards compared to those reported in Tirone and Ganguly (2010) by a factor of 2.8 at 700 °C, 2.2 at 800 °C and 1.8 at 900 °C, independent of cooling rate. This difference arises mainly because the values of D*Fe that are calculated at these temperatures using data from this work are higher than those calculated using Ganguly et al. (1998). 6.00 Ganguly and Tirone (1999) calculated closure temperature profiles in crystals surrounded by a homogeneous infinite matrix for different values of T0, M and crystal geometry. The results show that for T0 in the range of 700–1,100 °C, the temperatures calculated from the rim and matrix compositions are significantly lower than T0 if M [ 0.01. Using this value of M and the diffusion data presented in this paper, we have calculated the size of garnet grains below which the rim compositions would reflect significant resetting of temperature during cooling if these are surrounded by a homogeneous infinite matrix. The results are illustrated in Fig. 9. The basic observations from Figs. 8 and 9 are that (i) slow cooling (*1–2 °C/my) from temperatures[750 °C would reset the cores of several-mm-sized garnet grains that are enclosed in a medium where diffusion rates are infinitely fast (e.g., an infinite mass of biotite, rather than pyroxenes), (ii) rapid cooling ([100 °C/my) would almost always retain compositions from metamorphic peak at cores of crystals. On the other hand, it would be almost impossible to not reset the rim compositions of garnets cooling from temperatures[750 °C unless cooling is faster than several 1,000 °C/my (Fig. 9). For medium-grade metamorphic rocks (500–600 °C), the rim compositions are likely to retain (Fig. 9) the peak metamorphic compositions unless cooling is exceedingly slow. Thus, these diagrams may be used to estimate whether rim or core compositions should be used to determine peak 6.00 5.00 5.00 r /My 5 10 20 3.00 20 20 10 2.00 100 10 0 10 1.00 2 °C 5 2 5 2.00 Radius (mm) 4.00 3.00 2° C/M yr R a d iu s ( m m ) 4.00 100 1.00 0.00 650 700 750 800 850 T0 (°C) 0.00 Fig. 8 Radius of spherical grains that would retain their core compositions without resetting, for cooling from different peak temperatures, at different rates. The calculations are shown for binary chemical diffusion in a garnet of composition XFe = 0.75, XMg = 0.25. Bold solid lines show results obtained using diffusion coefficients reported in this work; fine solid lines show the results obtained by Tirone and Ganguly (2010) using data from Ganguly et al. (1998) for comparison 123 650 700 750 800 850 T0 (°C) Fig. 9 Curves showing grain sizes for which there would be significant resetting of rim compositions, for different peak temperatures and cooling rates. Calculations are for a chemical diffusion coefficient in a binary garnet of compositions XFe = 0.75, XMg = 0.25 (same as in Fig. 8) Contrib Mineral Petrol metamorphic temperatures at different grades. Curves such as the ones illustrated in Figs. 8 or 9 can be easily calculated using the Arrhenius parameters presented here and Eq. (10) for any arbitrary garnet size and cooling rate. Conclusions We have used numerical simulations and data from new experiments to clarify several aspects of multicomponent diffusion in garnets. Comparison of the compositions of experimental garnet diffusion couples and those commonly found in nature in metamorphic and mantle-derived rocks indicates that the natural garnet compositions, with the exception of some rather Ca-rich garnets from eclogites, are well represented by the experimental couples. Diffusion in Mn-rich, low-to-medium-grade pelitic garnets occur along different compositional vectors and have a different behavior (D*Fe*D*Mg). Therefore, these should be modeled using different diffusion coefficients from those used for modeling diffusion in more Mn-poor, Fe–Mg garnets (D*Fe \ D*Mg). In this study, we have focused on the diffusion behavior in garnets of the second kind. Incorporation of the effects of thermodynamic nonideality in the diffusion models alters the shapes of diffusion profiles. Therefore, use of non-ideal models helps to obtain better numerical fits to measured concentration profiles. Calculated diffusion penetration distances (or equivalently, retrieved diffusion coefficients from a given profile length) differ significantly between ideal and nonideal models for some unusual Mg–Mn–Ca-rich garnets only. Accounting for these effects, plus a more generalized convolution correction to profile shapes, allows us to obtain better defined diffusion coefficients from experimental runs. We have developed and tested a numerical method to calculate best-fit diffusion coefficients from experimentally induced concentration profiles. This method yields diffusion coefficients that are similar to the visual fitting method that has been used until now. The latter has the advantage that a larger range of diffusion couple compositions can be analyzed using the method because effects of compositional dependence of diffusion coefficients and asymmetry of profile shapes, which occur when the difference in composition between the two ends of the diffusion couple is large, can be handled. Here, we present data obtained using the objective numerical fitting method. The numerical calculations, combined with a theoretical analysis, further revealed that there are limitations to the application of these methods to extract multiple diffusion coefficients from a single run. When one of the diffusion coefficients becomes much faster or slower than the rest, or when the diffusion couple has a composition that is dominated by one component ([75 %), then profile shapes become insensitive to one or more diffusion coefficients and these cannot be retrieved by modeling the concentration profiles. We have carried out four new diffusion couple experiments at pressures of 25 kbar and at temperatures between 1,260 and 1,400 °C. Results from these experiments were combined with earlier results obtained using a similar experimental setup to obtain better constrained Arrhenius parameters for diffusion in Fe–Mg garnets. Consistency of results obtained in the new experiments that used a different population of garnets with the older results indicates that (a) for primarily Fe–Mg solid solutions, it is possible to use a single set of Arrhenius parameters to describe the diffusion behavior reasonably well and (b) trace element compositions of garnets do not affect the diffusion coefficients of Fe and Mg substantially. It is found that D*Fe \ D*Mg not only at experimental conditions, but also on extrapolation down to lower temperatures. The diffusion rates obtained using these parameters indicate that (a) it may be difficult to preserve UHT conditions in cores of garnets that are enclosed in a matrix where diffusion is infinitely fast, (b) rims of high-grade garnets will always be reset and (c) rims of low-to-medium-grade garnets will record peak temperature conditions at their rims. Acknowledgments We thank the German Science Foundation (DFG) for generously supporting this work. Thanks are due to J. Van Orman and an anonymous reviewer for constructive reviews. SAB was supported by the SFB 526 Program of the German Science Foundation and an INSA-DFG Fellowship funded the visit of SKB to Bochum. JG gratefully acknowledges the support from Alexander Humboldt foundation revisit program and US National Science Foundation grant No. EAR-1016189 for his participation in this project. References Carlson WD (2002) Scales of equilibrium and rates of equilibration during metamorphism. Am Mineral 87:185–204 Carlson WD (2006) Rates of Fe, Mg, Mn and Ca diffusion in garnet. Am Mineral 91:1–11 Chakraborty S, Ganguly J (1991) Compositional zoning and cation diffusion in garnets. In: Ganguly J (ed) Diffusion, atomic ordering and mass transport. Advances in physical geochemistry, vol 8. Springer, New York, pp 120–175 Chakraborty S, Ganguly J (1992) Cation diffusion in aluminosilicate garnets: experimental determination in spessartine-almandine diffusion couples, evaluation of effective binary diffusion coefficients, and applications. Contrib Mineral Petrol 111:74–86 Chakraborty S, Rubie DC (1996) Mg tracer diffusion in aluminosilicate garnets at 750–850 °C, 1 atm and 1300 °C, 8.5 GPa. Contrib Mineral Petrol 122:406–414 Cheng W, Ganguly J (1994) Some aspects of multicomponent excess free energy models with subregular binaries. Geochim Cosmochim Acta 58:3763–3767 123 Contrib Mineral Petrol Cygan RT, Lasaga AC (1985) Self-diffusion of magnesium in garnet at 750 to 900 °C. Am J Sci 285:328–350 Dodson MH (1973) Closure temperature in cooling geochronological and petrological systems. Contrib Mineral Petrol 40:259–274 Elphick SC, Ganguly J, Loomis TP (1985) Experimental determination of cation diffusivities in aluminosilicate garnets I. Experimental methods and interdiffusion data. Contrib Mineral Petrol 90:36–44 Ganguly (2010) Cation diffusion kinetics in garnet and geological applications. In: Zhang Y, Cherniak D (eds) Diffusion in minerals and melts. Reviews in mineralogy and geochemistry, vol 72. Mineralogical Society of America, pp 559–601 Ganguly J, Saxena SK (1987) Mixtures and mineral reactions. Springer, Berlin Ganguly J, Tirone M (1999) Diffusion closure temperature and age of a mineral with arbitrary extent of diffusion: Theoretical formulation and applications. Earth Planet Sci Lett 170:131–140 Ganguly J, Bhattacharya RN, Chakraborty S (1988) Convolution effect in the determination of compositional profiles and diffusion coefficients by microprobe step scans. Am Mineral 73:901–909 Ganguly J, Cheng W, Tirone M (1996) Thermodynamics of aluminosilicate garnet solid solution: new experimental data, an optimized model and thermometric applications. Contrib Mineral Petrol 126:137–151 Ganguly J, Cheng W, Chakraborty S (1998) Cation diffusion in aluminosilicate garnets: experimental determination in pyropealmandine diffusion couples. Contrib Mineral Petrol 131:171–180 Holzapfel C, Chakraborty S, Rubie DC, Frost DJ (2007) Effect of pressure on Fe–Mg, Ni and Mn diffusion in (FexMg1-x)2SiO4 olivine. Phys Earth Planet Inter 162:186–198 Kohn MJ (2003) Geochemical zoning in metamorphic minerals. In: Rudnick R (ed) Treatise on geochemistry, vol 3. The crust. Elsevier, pp 229–261 123 Lasaga AC (1979) Multicomponent exchange and diffusion in silicates. Geochim Cosmochim Acta 43:455–468 Loomis TP, Ganguly J, Elphick SC (1985) Experimental determination of cation diffusivities in aluminosilicate garnets II. Multicomponent simulation and tracer diffusion coefficients. Contrib Mineral Petrol 90:45–51 Muggianu YM, Gambino M, Bros JP (1975) Enthalpies of formation of liquid alloy bismuth-gallium-tin at 723 K. Choice of an analytical representation of integral and partial excess functions of mixing. J Chim Phys 72:83–88 Nelder JA, Mead R (1965) A simplex method for function minimization. Comput J 7:308–313 Perchuk AL, Burchard M, Schertl HP, Maresch WV, Gerya TV, Bernhardt HJ, Vidal O (2009) Diffusion of divalent cations in garnet: multi-couple experiments. Contrib Mineral Petrol 157:573–592 Sano J, Ganguly J, Hervig R, Dohmen R, Zhang X (2011) Neodymium diffusion in orthopyroxene: experimental studies and applications to geological and planetary problems. Geochim Cosmochim Acta 75:4684–4698 Schwandt CS, Cygan RT, Westrich HR (1995) Mg self-diffusion in pyrope garnet. Am Mineral 80:483–490 Smigelskas AD, Kirkendall EO (1947) Zinc diffusion in alpha brass. Trans AIME 171:130–142 Tirone M, Ganguly J (2010) Garnet compositions as recorders of P– T–t history of metamorphic rocks. Gondwana Res 18:138–146 Tirone M, Ganguly J, Dohmen R, Langenhorst F, Herwig R, Becker H-W (2005) Rare earth diffusion kinetics in garnet: experimental studies and applications. Geochim Cosmochim Acta 69:2385–2398 Vielzeuf D, Baronnet A, Perchuk AL, Laporte D, Baker MB (2007) Calcium diffusivity in alumino-silicate garnets: an experimental and ATEM study. Contrib Mineral Petrol 154:153–170 Wohl K (1953) Thermodynamic evaluation of binary and ternary liquid systems. Chem Eng Progr 49:8–219